Is Vanadate Reduced by Thiols under Biological Conditions

Apr 1, 2010 - Department of Chemistry, Colorado State University, Fort Collins, Colorado 80523-1872. ‡ Biochemistry Department, State University of ...
0 downloads 4 Views 1010KB Size
Inorg. Chem. 2010, 49, 4245–4256 4245 DOI: 10.1021/ic100080k

)

Is Vanadate Reduced by Thiols under Biological Conditions? Changing the Redox Potential of V(V)/V(IV) by Complexation in Aqueous Solution Debbie C. Crans,*,† Boyan Zhang,† Ernestas Gaidamauskas,† Anastasios D. Keramidas,†, Gail R. Willsky,‡ and Chris R. Roberts§ Department of Chemistry, Colorado State University, Fort Collins, Colorado 80523-1872, ‡Biochemistry Department, State University of New York at Buffalo, School of Medicine and Biomedical Sciences, Buffalo, New York 14214, and §Roche Colorado Corporation, 2075 North 55th Street, Boulder, Colorado 80301. Current address: Dept. of Chemistry, University of Cyprus, 1678 Nicosia, Cyprus

)



Received January 14, 2010

Although dogma states that vanadate is readily reduced by glutathione, cysteine, and other thiols, there are several examples documenting that vanadium(V)-sulfur complexes can form and be observed. This conundrum has impacted life scientists for more than two decades. Investigation of this problem requires an understanding of both the complexes that form from vanadium(IV) and (V) and a representative thiol in aqueous solution. The reactions of vanadate and hydrated vanadyl cation with 2-mercaptoethanol have been investigated using multinuclear NMR, electron paramagnetic resonance (EPR), and UV-vis spectroscopy. Vanadate forms a stable complex of 2:2 stoichiometry with 2-mercaptoethanol at neutral and alkaline pH. In contrast, vanadate can oxidize 2-mercaptoethanol; this process is favored at low pH and high solute concentrations. The complex that forms between aqueous vanadium(IV) and 2-mercaptoethanol has a 1:2 stoichiometry and can be observed at high pH and high 2-mercaptoethanol concentration. The solution structures have been deduced based on coordination induced chemical shifts and speciation diagrams prepared. This work demonstrates that both vanadium(IV) and (V)-thiol complexes form and that redox chemistry also takes place. Whether reduction of vanadate takes place is governed by a combination of parameters: pH, solute- and vanadate-concentrations and the presence of other complexing ligands. On the basis of these results it is now possible to understand the distribution of vanadium in oxidation states (IV) and (V) in the presence of glutathione, cysteine, and other thiols and begin to evaluate the forms of the vanadium compounds that exert a particular biological effect including the insulin-enhancing agents, antiamoebic agents, and interactions with vanadium binding proteins. Introduction The presence of oxidants and reductants in solution usually leads to redox chemistry.1 A range of vanadium-sulfur complexes and their chemistries have demonstrated that vanadiumsulfur complexes have interest as models for nitrogenase *To whom correspondence should be addressed. E-mail: crans@ lamar.colostate.edu. (1) Preliminary presentation was presented in: Crans, D. C.; Zhang, B. Y.; Gaidamauskas, E.; Keramidas, A.; Roberts, C. Abstracts of Papers, 223rd National Meeting of the American Chemical Society, Orlando, FL, April 7-11, 2002; American Chemical Society: Washington, DC, 2002; INOR-242. (2) Hsu, H. F.; Su, C. L.; Gopal, N. O.; Wu, C. C.; Chu, W. C.; Tsai, Y. F.; Chang, Y. H.; Liu, Y. H.; Kuo, T. S.; Ke, S. C. Eur. J. Inorg. Chem. 2006, 1161–1167. (3) Kawakami, N.; Ueki, T.; Amata, Y.; Kanamori, K.; Matsuo, K.; Gekko, K.; Michibata, H. Biochim. Biophys. Acta, Proteins Proteomics 2009, 1794, 674–679. (4) Ueki, T.; Shintaku, K.; Yonekawa, Y.; Takatsu, N.; Yamada, H.; Hamada, T.; Hirota, H.; Michibata, H. Biochim. Biophys. Acta, Gen. Subj. 2007, 1770, 951–957. (5) Singh, S.; Bharti, N.; Mohapatra, P. P. Chem. Rev. 2009, 109, 1900– 1947. (6) Maurya, M. R.; Kumar, A.; Bhat, A. R.; Azam, A.; Bader, C.; Rehder, D. Inorg. Chem. 2006, 45, 1260–1269.

r 2010 American Chemical Society

chemistry,2 as the vanadium binding proteins,3,4 as antiamoebic agents,5-7 as antidiabetic agents,8 and as catalysts.9,10 The literature provides many examples that rapid reduction of vanadium(V) compounds takes place in the presence of biological thiols such as glutathione, cysteine, 2-mercaptoethanol, and dithiothreitol (DTT) under physiological conditions.11-17 Redox reactions of glutathione and cysteine (7) Maurya, M. R.; Khurana, S.; Shailendra; Azam, A.; Zhang, W. J.; Rehder, D. Eur. J. Inorg. Chem. 2003, 1966–1973. (8) Bolechala, K.; Zborowski, K. K.; Zuchowski, G.; Proniewicz, L. M. J. Phys. Org. Chem. 2009, 22, 994–1002. (9) Maurya, M. R.; Kumar, U.; Correia, I.; Adao, P.; Pessoa, J. C. Eur. J. Inorg. Chem. 2008, 577–587. (10) Chu, W. C.; Wu, C. C.; Hsu, H. F. Inorg. Chem. 2006, 45, 3164–3166. (11) Crans, D. C.; Smee, J. J.; Gaidamauskas, E.; Yang, L. Q. Chem. Rev. 2004, 104, 849–902. (12) Cantley, L. C.; Aisen, P. J. Biol. Chem. 1979, 254, 1781–1784. (13) Sakurai, H.; Shimomura, S.; Ishizu, K. Inorg. Chim. Acta 1981, 55, L67–L69. (14) Willsky, G. R.; White, D. A.; Mccabe, B. C. J. Biol. Chem. 1984, 259, 3273–3281. (15) Li, J. P.; Elberg, G.; Crans, D. C.; Shechter, Y. Biochemistry 1996, 35, 8314–8318. (16) Legrum, W. Toxicology 1986, 42, 281–289. (17) Shi, X. L.; Sun, X. Y.; Dalal, N. S. FEBS Lett. 1990, 271, 185–188.

Published on Web 04/01/2010

pubs.acs.org/IC

4246 Inorganic Chemistry, Vol. 49, No. 9, 2010 have also been reported with chromium(VI),18,19 iron(III),20,21 and ruthenium(III)22,23 compounds. Reports on formation and characterization of stable vanadium(V) complexes with thiol-containing ligands2,5-7,24-26 and their solution chemistry show26-28 that these complexes can be stable. Slow reduction of vanadium(V) by thiols in biological systems15,29 implies that vanadium(V) complexes can form, be characterized, and possibly isolated.24-28,30,31 It was our objective in this study to investigate the seemingly contradictory reports by exhaustively characterizing the reactions of vanadium(V) (oxidant) with 2-mercaptoethanol (reductant), as well as the solution chemistry of both the vanadium(V) and (IV) products. This work provides guidance reconciling this controversy. Vanadium salts and other vanadium coordination compounds induce insulin-enhancing effects in vitro and in vivo. The human studies with vanadyl sulfate and sodium metavanadate show that simple vanadium salts have reasonable effects even at doses significantly below the “efficacious” dose.32-34 Studies with bis(maltolato)-oxovanadium(IV),35-37 bis(acetylacetonato)-oxovanadium(IV),38 and many other oxovanadium(IV) complexes39-42 demonstrate that these coordination complexes have higher potency than simple vanadium salts. Although the mechanism of action of vanadium compounds remains elusive, it is generally accepted that inhibition of protein tyrosine phosphatase(s), and subsequent (18) Levina, A.; Lay, P. A. Inorg. Chem. 2004, 43, 324–335. (19) Brauer, S. L.; Wetterhahn, K. E. J. Am. Chem. Soc. 1991, 113, 3001– 3007. (20) Valko, M.; Rhodes, C. J.; Moncol, J.; Izakovic, M.; Mazur, M. Chem.-Biol. Interact. 2006, 160, 1–40. (21) Sisley, M. J.; Jordan, R. B. Inorg. Chem. 1995, 34, 6015–6023. (22) Frasca, D. R.; Clarke, M. J. J. Am. Chem. Soc. 1999, 121, 8523–8532. (23) Wang, F. Y.; Weidt, S.; Xu, J. J.; Mackay, C. L.; Langridge-Smith, P. R. R.; Sadler, P. J. J. Am. Soc. Mass. Spectrom. 2008, 19, 544–549. (24) Farahbakhsh, M.; Nekola, H.; Schmidt, H.; Rehder, D. Chem. Ber.-Rec. 1997, 130, 1129–1133. (25) Cornman, C. R.; Stauffer, T. C.; Boyle, P. D. J. Am. Chem. Soc. 1997, 119, 5986–5987. (26) Bhattacharyya, S.; Batchelor, R. J.; Einstein, F. W. B.; Tracey, A. S. Can. J. Chem. 1999, 77, 2088–2094. (27) Paul, P. C.; Tracey, A. S. J. Biol. Inorg. Chem. 1997, 2, 644–651. (28) Crans, D. C.; Mahroof Tahir, M.; Keramidas, A. D. Mol. Cell. Biochem. 1995, 153, 17–24. (29) Chasteen, N. D.; Grady, J. K.; Holloway, C. E. Inorg. Chem. 1986, 25, 2754–2760. (30) Tsagkalidis, W.; Rodewald, D.; Rehder, D. Inorg. Chem. 1995, 34, 1943–1945. (31) Nanda, K. K.; Sinn, E.; Addison, A. W. Inorg. Chem. 1996, 35, 1–2. (32) Goldfine, A. B.; Simonson, D. C.; Folli, F.; Patti, E.; Kahn, C. R. Mol. Cell. Biochem. 1995, 153, 217–231. (33) Willsky, G. R.; Goldfine, A. B.; Kostyniak, P. J.; McNeill, J. H.; Yang, L. Q.; Khan, H. R.; Crans, D. C. J. Inorg. Biochem. 2001, 85, 33–42. (34) Cohen, N.; Halberstam, M.; Shlimovich, P.; Chang, C. J.; Shamoon, H.; Rossetti, L. J. Clin. Invest. 1995, 95, 2501–2509. (35) Thompson, K. H.; McNeill, J. H.; Orvig, C. Chem. Rev. 1999, 99, 2561–2571. (36) Mcneill, J. H.; Yuen, V. G.; Hoveyda, H. R.; Orvig, C. J. Med. Chem. 1992, 35, 1489–1491. (37) Orvig, C.; McNeill, J. H.; Vasilevskis, J. In Metals in Biology: Vanadium; Sigel, H., Sigel, A., Eds.; Marcel Dekker: New York, 1995; Vol. 31, pp 575-594. (38) Reul, B. A.; Amin, S. S.; Buchet, J. P.; Ongemba, L. N.; Crans, D. C.; Brichard, S. M. Br. J. Pharmacol. 1999, 126, 467–477. (39) Kiss, T.; Jakusch, T.; Hollender, D.; Dornyei, A.; Enyedy, E. A.; Pessoa, J. C.; Sakurai, H.; Sanz-Medel, A. Coord. Chem. Rev. 2008, 252, 1153–1162. (40) Adachi, Y.; Yoshida, J.; Kodera, Y.; Katoh, A.; Takada, J.; Sakurai, H. J. Med. Chem. 2006, 49, 3251–3256. (41) Sakurai, H.; Fujii, K.; Fujimoto, S.; Fujisawa, Y.; Takechi, K.; Yasui, H. ACS Symp. Ser. 1998, 711, 344–352. (42) Sakurai, H.; Fujii, K.; Watanabe, H.; Tamura, H. Biochem. Biophys. Res. Commun. 1995, 214, 1095–1101.

Crans et al. 0

Table 1. Standard Apparent Reduction Potentials (E 0 ) of Vanadate and Thiols at 298 K and pH =7.0 or pD = 7.0a 0

redox couple H2VO4- þ 4Hþ þ e- a VO2þ þ 3H2O 2H2VO4- þ 4Hþ þ 2e- a V2O4 þ 4H2O 1/2 RSSR þ Hþ þ e- a RSH R = HOCH2CH2 (2-Mercaptoethanol) R = H2NCH(CO2H)CH2CH2CONHCH(CONHCH2CO2H)CH2 (Glutathione) R = HOOC(NH2)CHCH2 (Cysteine) R = 1/2 H2C(CHOH)2CH2 (Dithiothreitol)

E 0 /V vs SHE (pH 7 or pD 7) -0.34149 0.19549 -0.253a,51; -0.260a,52 -0.252a,52; -0.262a,51 -0.2250 -0.245a,51 -0.327a,52

a Standard apparent reduction potentials determined from thioldisulfide equilibrations in D2O at pD = 7.0 and 298 K.

activation of a protein kinase(s), is involved.15,43-46 The different modes of action of vanadate and peroxovanadate with protein tyrosine phosphatases (1B and Leukocyte Antigen Related (LAR)) have been characterized, the former acting as a reversible inhibitor and the latter as a redox active irreversible inhibitor.46,47 Mechanistic studies are complicated by the rich hydrolytic chemistry of vanadium compounds and by the facile interconversion between oxidation states (especially between (V) and (IV)) under physiological conditions. Despite (or perhaps because of) the rapid interconversion between the two oxidation states, sufficient information to quantitatively evaluate both vanadium(V) and vanadium(IV) chemistry with one thiol-containing ligand has not previously been reported. The half reactions and standard apparent reduction potentials at pH = 7.0 for the systems relevant to this study are summarized in Table 1. On the basis of the values of the standard reduction potential for one electron reduction, thiols (E ∼ -0.26 V) should not be able to reduce vanadate (E = -0.341 V). Since this reaction occurs, it is clear that concentrations of reaction components and pH play a critical role in whether redox reactions take place. In the case of thiols, the observed solution potential will be similar to the formal standard potential (-0.25 ( 0.01 V in case of 2-mercaptoethanol). This is not the case with the vanadium(IV)/vanadium(V) redox couple because of the nature of the aqueous vanadium chemistry at neutral pH.48 At neutral pH only low concentrations of monomeric vanadyl cation (∼10-8 M) exist because of the low solubility product of vanadyl hydroxide. Furthermore, only low concentrations of vanadate monomer (∼10-3 M) exist because of the high stability of vanadium(V) oligomers. In a solution containing 1 mM H2VO4- the reduction potential will equal 0.01 V (43) Li, M.; Ding, W. J.; Baruah, B.; Crans, D. C.; Wang, R. L. J. Inorg. Biochem. 2008, 102, 1846–1853. (44) Yuan, C. X.; Lu, L. P.; Gao, X. L.; Wu, Y. B.; Guo, M. L.; Li, Y.; Fu, X. Q.; Zhu, M. L. J. Biol. Inorg. Chem. 2009, 14, 841–851. (45) Vardatsikos, G.; Mehdi, M. Z.; Srivastava, A. K. Int. J. Mol. Med. 2009, 24, 303–309. (46) Shechter, Y.; Elberg, G.; Shisheva, A.; Gefel, D.; Sekar, N.; Qian, S.; Bruck, R.; Gershonov, E.; Crans, D. C.; Goldwasser, Y.; Fridkin, M.; Li, J. ACS Symp. Ser. 1998, 711, 308–315. (47) Huyer, G.; Liu, S.; Kelly, J.; Moffat, J.; Payette, P.; Kennedy, B.; Tsaprailis, G.; Gresser, M. J.; Ramachandran, C. J. Biol. Inorg. Chem. 1997, 272, 843–851. (48) The studies described in this work are at the millimolar concentration of vanadium salt and complexes. A solution potential could be observed as its standard value of -0.34 V when both concentrations of H2VO4- and VO2þ are kept equal (∼10-8 M). The considerations described here therefore do not apply to systems of such low vanadium concentration and lack of complex formation (which never would exist in any biological system).

Article

when the vanadyl cation concentration is determined by the solubility product of [VO(OH)2]. This suggests that thiols are readily oxidized by vanadate. However, the reaction quotient depends not only on the redox potentials describing reduction of vanadate and oxidation of thiol but also on the oligomerization and complex formation. All equilibria must be considered in the overall analysis of complex formation and redox chemistry. In the absence of complexation between vanadium(V) and reducing thiol, redox reactions are thermodynamically favored when a large excess of vanadium(V) over vanadium(IV) is present. Formation of vanadium(V)-thiol complexes can prevent redox chemistry, if the vanadium(V) complex is stabilized more than the vanadium(IV) complex. We have characterized the stoichiometry, stability, and structure of the species that form in solution with 2-mercaptoethanol, a common reducing component in enzyme assays. The complexation chemistry of vanadate with 1,2-diols has been characterized both in solution and in the solid state.53-55 While this work was underway, a crystal structure of [(VO2)2(OCH2CH2S)2]- and some aqueous and organic solution properties were reported, and are used as structural benchmark data.26 In this work, we study the interplay between complexes of different oxidation states and define the conditions under which vanadate and/or vanadyl cation form complexes with 2-mercaptoethanol. In aqueous solution, the vanadium(V)-thiolate complex (major 2:2 complex) is surprisingly stable. When the vanadium(V) is reduced, a vanadium(IV) complex with a 1:2 stoichiometry forms, attesting to the dramatic influence oxidation state has on complex stoichiometry and structure. Understanding the vanadium(V)-thiolate complexes and their redox chemistry is essential to understand the biological effects of vanadium compounds. Experimental Section Chemicals. Reagent grade compounds were purchased from Aldrich Chemical Co, Fisher Scientific, Mallinckrodt, and from Acros Organics (2-hydroxyethyl disulfide (95%)) and used without further purification. Deuterium oxide (Cambridge Isotope Laboratories, 99.9%), NaOD (Merck & Co., 98 atom % D) and DCl (Aldrich, 99.5 atom % D) and [17O]H2O (ICON, 10.43 atom % 17O) were used as received. Solutions. Deionized water (17.9 MΩ 3 cm) was used to prepare all the H2O solutions. Stock solutions of sodium vanadate in D2O or in H2O were prepared by adding 1 equiv of V2O5 to 2 equiv of 2 M NaOD or NaOH. The suspension was heated until the solid was completely dissolved. After cooling to ambient temperature the solution was further diluted with D2O or H2O to the desired vanadium concentration (100 or 25 mM). When stock solution with lower pH was needed (Corning 140 pH meter equipped with a combination microelectrode), the (49) Pourbaix, M. In Atlas of electrochemical equilibria in aqueous solutions, 1st English ed.; Pergamon Press: Oxford, NY, 1966; pp 234-245. (50) Reipa, V. Bioelectrochemistry 2004, 65, 47–49. (51) Millis, K. K.; Weaver, K. H.; Rabenstein, D. L. J. Org. Chem. 1993, 58, 4144–4146. (52) Lees, W. J.; Whitesides, G. M. J. Org. Chem. 1993, 58, 642–647. (53) Ray, W. J.; Crans, D. C.; Zheng, J.; Burgner, J. W.; Deng, H.; Mahrooftahir, M. J. Am. Chem. Soc. 1995, 117, 6015–6026. (54) Angusdunne, S. J.; Batchelor, R. J.; Tracey, A. S.; Einstein, F. W. B. J. Am. Chem. Soc. 1995, 117, 5292–5296. (55) Zhang, B. Y.; Zhang, S. W.; Wang, K. Chem. Soc., Dalton Trans. 1996, 3257–3263. (56) The pH of the solution was adjusted to 7.8 by the addition of HCl or DCl. The pH values reported here are not corrected for deuterium isotope effects although they were measured in the presence of D2O.

Inorganic Chemistry, Vol. 49, No. 9, 2010

4247

addition of DCl resulted in a bright yellow solution, which was then heated until the yellow color disappeared and the pH remained constant at room temperature (300 ( 1 K).56 The series of 100 mM solutions of 2-mercaptoethanol in D2O at various pH values were prepared, under N2, just before use. Borate buffer stock solution (∼0.7 M) was prepared by dissolving of H3BO3 in D2O (or H2O) and adjusting pH with NaOD (or NaOH) solution as needed. Mixed aqueous-acetonitrile solutions of vanadate and 2-mercaptoethanol were prepared by adding an aqueous (D2O) vanadate solution to a 2-mercaptoethanol solution in acetonitrile. The solutions for spectroscopic studies were purged with N2 for 15 min to prevent air oxidation of the thiol and kept under a N2 atmosphere at 300 ( 1 K. Studies without buffers were monitored carefully, and no more than a 0.5 pH unit change was found over any 4 h experiment. All the spectroscopic measurements were made no more than 1 h after mixing the stock vanadate and thiol solutions. NMR Spectroscopy. NMR spectra were recorded on a Bruker ACP-300 spectrometer operating at room temperature (300 ( 1 K) at 79 MHz for 51V, 300 MHz for 1H, 76 MHz for 13C, and 41 MHz for 17O. 51V NMR spectra were recorded with an acquisition time of 0.04 s, no relaxation delay, 60 pulse-width, and 50 kHz spectral width. Chemical shifts are reported relative to external references of VOCl3 0 ppm for 51V, TMS for 1H and 13 C, and H2O for 17O. Samples for [17O] NMR spectroscopy were enriched with [17O]H2O to 3% [17O]H2O. Parameters for an initial 17O NMR spectrum with a 65,000 Hz spectral window, 0.10 s acquisition time, 90 pulse width, and no relaxation delay preceded a second acquisition using a narrower spectral window of ∼5,000 Hz resetting the transmitter frequency as necessary. Total collection time per spectrum was about 30 min. Baseline corrections were applied. Quantitative analyses of all the 1H and 51V NMR spectra were performed using Bruker integration software. The concentrations of the vanadium and thiol species in each sample were calculated on the basis of the observed molar fraction of 1H and 51 V signals, and the total added vanadium and thiol concentrations assuming reduction to the vanadium(IV) species was negligible. This assumption was confirmed by recording electron paramagnetic resonance (EPR) spectra of solutions with similar concentrations of thiol and vanadate where no more than 2% of a vanadium(IV) species was observed. EPR Spectroscopy. EPR spectra were recorded using 1 mm o.d. capillary inserted in a 4 mm o.d. quartz EPR tube for ambient temperature (300 ( 1 K) spectra and 4 mm o.d. quartz tubes for low temperature (140 ( 1 K) spectra, on a Bruker ESP 300 spectrometer. To acquire ambient temperature spectra, the spectrometer was operated at 9.77 GHz in the X-band with a microwave power of 20 mW. Previous studies57 have documented that these parameters avoid signal saturation in similar systems. To acquire the low temperature spectra, the spectrometer was operated at 9.465 GHz in the X-band with a microwave power of 2 mW. A modulation frequency of 100 kHz, a modulation amplitude of 7.95 G, a time constant of 20.48 ms, and a conversion time of 81.92 ms were employed at low temperature. The spectrometer was calibrated with 1,1-diphenyl-2-picrylhydrazyl (DPPH, g = 2.0037).57 Computer simulations of the fluid solution EPR spectra were performed with ASYM, which includes second-order corrections to the nuclear hyperfine interaction58 and the dependence of the line width on vanadium nuclear spin.59,60 The estimated (57) Amin, S. S.; Cryer, K.; Zhang, B. Y.; Dutta, S. K.; Eaton, S. S.; Anderson, O. P.; Miller, S. M.; Reul, B. A.; Brichard, S. M.; Crans, D. C. Inorg. Chem. 2000, 39, 406–416 and references therein. (58) Atherton, N. M. In The Isotropic Hyperfine Interaction; Atherton, N. M., Ed.; Horwood/Prentice Hall: New York, 1993; pp 114-122. (59) Wilson, R.; Kivelson, D. J. Chem. Phys. 1966, 44, 154–168. (60) Chasteen, N. D.; Hanna, M. W. J. Phys. Chem. 1972, 76, 3951–3958.

4248 Inorganic Chemistry, Vol. 49, No. 9, 2010

Crans et al.

uncertainties are (based on data collection in duplicate and triplicate on identical samples, as well as simulation of one set of spectral data): hyperfine couplings, ( 110-4 cm-1; g values, ( 0.002; and concentration of species, ( 2%. Concentrations of vanadium(IV) species were determined by double-integration of the first-derivative EPR signals and calculated based on a calibration curve generated from solutions of 0.045, 0.091, 0.45, 0.91, 1.8, 4.83, 6.0, and 10.0 mM VOSO4 in 0.1 M H2SO4. Unless otherwise specified all spectra were recorded within 30 min of solution preparation. UV-visible Spectroscopy. Absorption spectra were recorded on a Perkin-Elmer Lambda 4B spectrometer equipped with a HAAKE A81 circulating temperature bath (300.2 ( 0.1 K). Kinetic Measurements. The rate measurements for the formation of disulfide and other species in solutions containing vanadate and 2-mercaptoethanol (from 50 to 300 mM) were done by 1H and 51V NMR spectroscopy. Each measurement was conducted in duplicate. Studies at pH 8.95 were carried out in 0.40 M KCl in 0.40 M borate buffer. All anaerobic samples were purged with nitrogen for 15 min before the solutions were mixed, then the solution pH was adjusted and the sample placed in a NMR tube under nitrogen atmosphere. A control sample was prepared for comparison, containing all reagents except vanadate. During the time of these experiments no more than 15% of the total 2-mercaptoethanol was oxidized to the disulfide. Data Analysis. Data analysis was carried out using an Excel spreadsheet. All deviations in parameters given indicate the 95% confidence level. The error bars on the figures represent the 95% confidence limit of at least three independent measurements. Speciation and Calculation of Formation Constants by NMR Spectroscopy. Vanadate and 2-mercaptoethanol deprotonate according to eqs 1 and 2, respectively, in the pH range 7.0 to 10.5. Given the dependence of the pKa values on ionic strength, for vanadate61,62 pKa values were determined at both the ionic strengths used in this study. At low ionic strength (no KCl) the deprotonation constant of 2-mercaptoethanol is 9.7 ( 0.1 (determined by both 1H NMR spectroscopy and potentiometric titration), and the second pKa of vanadate is 8.9 ( 0.2, consistent with those previously reported.61,63 At high ionic strength the pKa of 2-mercaptoethanol did not change (9.7), but the second pKa of vanadate decreased to 8.5 ( 0.2. H2 VO4 - a HVO4 2 - þ Hþ

ð1Þ

HOCH2 CH2 SH a HOCH2 CH2 S - þ Hþ

ð2Þ

Vanadate, as previously described,64 forms a series of oligomers including monomer (V1), dimer (V2), tetramer (V4), and pentamer (V5) that were measured by 51V NMR spectroscopy. The concentrations of 2-mercaptoethanol and the oxidation product 2-hydroxyethyl disulfide were calculated from the integration of 1H NMR spectra. Vanadate-2-mercaptoethanol complexes were determined using both 1H and 51V NMR spectra. Using the pKa values for H2VO4- and HOCH2CH2SH measured above, the concentrations of H2VO4- (or HVO42-) and HOCH2CH2SH (or HOCH2CH2S-) can be calculated from eqs 3 and 4. Here [V1] and [HOCH2CH2SH]T are the total concentrations of free vanadate monomer and [HOCH2CH2SH] þ [HOCH2CH2S-]. (61) Tracey, A. S.; Jaswal, J. S.; Angusdunne, S. J. Inorg. Chem. 1995, 34, 5680–5685. (62) Gresser, M. J.; Tracey, A. S. J. Am. Chem. Soc. 1985, 107, 4215–4220. (63) Townshend, A. Dictionary of analytical reagents, 1st ed.; Chapman & Hall: London; New York, 1993. (64) Crans, D. C. Comments Inorg. Chem. 1994, 16, 1–33.

½H2 VO4-  ¼

½V1  1 þ K a, H2 VO4- =½Hþ 

½HOCH2 CH2 SH ¼

½HOCH2 CH2 SHT 1 þ K a, H2 L =½Hþ 

ð3Þ

ð4Þ

Speciation and Calculation of Formation Constants by EPR Spectroscopy. The reactions to be considered around neutral pH in these solutions and the logarithms of the respective formation constants and solubility product are summarized in eqs 5 to 9.65-68 The vanadyl cation as [VO(H2O)5]2þ is observed only below pH 3. When the pH is increased, two hydrolysis products form: a paramagnetic [VO(OH)(H2O)]þ species (eq 5) and an EPR silent dimer [(VO)2(OH)2]2þ (eq 6).65,69 At neutral pH the precipitation of [VO(OH)2] takes place (eq 7), which redissolves in basic media. The formation of insoluble hydroxides has previously been used to calculate the concentration of vanadyl cation in the neutral pH range.66 Two soluble VO2þ:OHstoichiometries are generally used in solution studies in the pH range from 6 to 13: the 2:5 stoichiometry (EPR silent species [(VO)2(OH)5-]n; eq 8 for n = 1), and the 1:3 stoichiometry (EPR active species [VO(OH)3-]n; eq 9 for n = 1).68-70 Speciation studies in dilute (1-10 mM) vanadium(IV) solutions provide results consistent with n = 1, which is commonly used.71-73 VO2þ þ H2 O a ½VOðOHÞþ þ Hþ

2VO2þ þ 2H2 O a ½ðVOÞ2 ðOHÞ2 2þ þ 2Hþ

VO2þ þ 2OH - a ½VOðOHÞ2 V 2VO2þ þ 5H2 O a ½ðVOÞ2 ðOHÞ5  - þ 5Hþ VO2þ þ 3H2 O a ½VOðOHÞ3  - þ 3Hþ

log β ¼ - 6:0

ð5Þ

log β ¼ - 6:9 ð6Þ

log K sp ¼ - 22:0

ð7Þ

log β ¼ - 22:0 ð8Þ

log β ¼ - 18:5

ð9Þ

Results and Discussion Reaction of Vanadate with 2-Mercaptoethanol in Aqueous Solution. Mixing colorless aqueous solutions of 2-mercaptoethanol and vanadate at neutral or slightly alkaline pH immediately generates yellow or orange mixtures both in the presence and absence of oxygen. The color intensity is dependent on the concentration of reactants and solution pH. For example, mixing 12.0 mM mercaptoethanol and 12.0 mM vanadate at pH 7.5 produces a yellow solution, which changes color very slowly (65) Henry, R. P.; Mitchell, P. C.; Prue, J. E. J. Chem. Soc., Dalton Trans. 1973, 1156–1159. (66) Chasteen, N. D. In Biological Magnetic Resonance; Berliner, L., Reuben, J., Eds.; Plenum: New York, 1981; Vol. 3, pp 53-119. (67) Vilas Boas, L. V.; Costa Pessoa, J. In Comprehensive Coordination Chemistry; Wilkinson, G., Gillard, R. D., McCleverty, J. A., Eds.; Pergamon: New York, 1987; Vol. 3, pp 453-583. (68) Komura, A.; Hayashi, M.; Imanaga, H. Bull. Chem. Soc. Jpn. 1977, 50, 2927–2931. (69) Francavilla, J.; Chasteen, N. D. Inorg. Chem. 1975, 14, 2860–2862. (70) Iannuzzi, M. M.; Rieger, P. H. Inorg. Chem. 1975, 14, 2895–2899. (71) Pessoa, J. C.; Boas, L. F. V.; Gillard, R. D. Polyhedron 1990, 9, 2101– 2125. (72) Pessoa, J. C.; Boas, L. F. V.; Gillard, R. D.; Lancashire, R. J. Polyhedron 1988, 7, 1245–1262. (73) Pessoa, J. C.; Luz, S. M.; Cavaco, I.; Gillard, R. D. Polyhedron 1994, 13, 3177–3198.

Article

Inorganic Chemistry, Vol. 49, No. 9, 2010

4249

Figure 2. pH dependence of complex [1] and vanadate monomer [V1] is shown. Data points were obtained by quantitative 1H and 51V NMR spectroscopy on solutions containing vanadate (4.0 mM) and 2-mercaptoethanol (12.0 mM) in the absence of KCl.

Solutions of 2-mercaptoethanol and vanadate generated H NMR spectra containing signals due to 2-mercaptoethanol, one vanadium(V)-2-mercaptoethanol complex (complex 1), and the oxidation product of 2-mercaptoethanol: 2-hydroxyethyl disulfide (Figure 1, bottom). The two signals for free 2-mercaptoethanol ligand are centered at 2.68 and 3.70 ppm (Figure 1, bottom). The signals at 2.90 and 3.87 ppm were readily assigned to 2-hydroxyethyl disulfide. The remaining two signals at 3.05 and 4.31 ppm are due to the major complex of vanadium(V) with 2-mercaptoethanol. The chemical shifts of the complex are consistent with the involvement of at least one S-donor in the coordination sphere of the vanadium atom. Stoichiometry and Formation Constants of the Two Vanadium(V)-2-Mercaptoethanol Complexes. For determining the stoichiometry and formation constant of complex 1, 51V and 1H NMR spectra were recorded of freshly prepared solutions containing various concentrations of vanadate (2.0 to 20.0 mM) and 2-mercaptoethanol (2.0 to 20.0 mM). The pH stability of this complex was examined in the pH region of 7.0 to 10.5 as shown in Figure 2 (for a solution containing 4.0 mM vanadate and 12.0 mM 2-mercaptoethanol). Spectra recorded outside this pH range lead to less reliable quantitative data because of the rapid reduction of the vanadium(V), rapid thiol oxidation and/or the low complex concentration. The linear correlation shown in Figure 3a indicates that the complex 1, at -362 ppm, has a stoichiometry of 2:2; other complex stoichiometries do not fit the data (Supporting Information, Figure S1). No protons are taken up or released during the reaction. Given the respective pKa values for H2VO4- and 2-mercaptoethanol and the pH range studied, formation of a complex with a charge of (2-), as shown in eq 10, is consistent with the species present in the solution, their pKa values, and the structure characterized in the solid state.26 An analogous linear relationship between [1] and [H2VO4-]2[HOCH2CH2SH]2 was observed at low ionic strength (Supporting Information, Figure S2). The overall (Hþindependent) formation constant for complex 1 at low ionic strength was calculated using eq 11 to be (3.7 ( 0.1) 3 108 M-3. The pH independent formation constant of complex 1 in 3.0 M KCl was found to be (1.2 ( 0.1) 3 108 M-3 (Figure 3a). The formation constant for complex 1 at high ionic strength is 3-fold smaller than that at low 1

Figure 1. 51V (top) and 1H (bottom) NMR spectra of solutions containing vanadate (12.0 mM) and 2-mercaptoethanol (12.0 mM) at pH 8.20 ( 0.05. The spectra were recorded at 300 ( 1 K.

to pale green and eventually to a colorless solution over the course of 2 weeks. A deep orange color forms immediately upon mixing 300 mM mercaptoethanol and 300 mM vanadate solutions at pH 9.0. No 51V NMR signal for decavanadate could be observed. This orange solution changed to dark green and to greenish-black over the course of a few hours. In contrast, mixtures above pH 10.0 remain colorless. A blue-black microcrystalline precipitate can be isolated from concentrated solutions that have turned greenish-black. The isolated blue-black material, upon dissolution in acidic solution, generated the blue vanadyl ion ([VO(H2O)5]2þ) with the characteristic eight line hyperfine pattern in the EPR spectrum. Elemental analysis and IR results confirmed that the blue-black microcrystals are vanadyl oxide VO2.74 These data clearly demonstrate that vanadate, in the presence of a thiol such as 2-mercaptoethanol, is reduced to vanadium(IV). At low concentrations of vanadate and thiol the yellow vanadium(V)-thiolate complexes were observed at neutral pH and persisted for some time. The 51V NMR spectrum of a yellow solution of 12.0 mM vanadate and 12.0 mM 2-mercaptoethanol is shown in Figure 1 (top). This spectrum contains a major signal at -362 ppm (complex 1) and a minor signal at -385 ppm (complex 2) in addition to signals due to the labile monomeric and oligomeric vanadate oxoanions. Two complexes are formed by vanadate and DTT at pH 7.527 but previously only a 2:2 vanadium(V) complex with 2-mercaptoethanol ([(VO2)2(OCH2CH2S)2]2-) at -362 ppm was reported.26 In our system the 51V NMR chemical shifts for both complexes are similar to those reported for vanadate and DTT mixtures.27 (74) Hewston, T. A.; Nadler, M. P. J. Solid State Chem. 1987, 71, 278– 281.

4250 Inorganic Chemistry, Vol. 49, No. 9, 2010

Crans et al.

Figure 3. Plots of complexes [1] (signal at -362 ppm) and [2] (signal at -385 ppm) as a function of [H2VO4-]2[H2L]2 (here H2L = HOCH2CH2SH) and [1]2, respectively. Data points were obtained by quantitative 1H and 51V NMR spectroscopy.

ionic strength as anticipated for a doubly charged dinuclear complex.61,64 2H2 VO4 - þ 2HOCH2 CH2 SH a ½ðVO2 Þ2 ðOCH2 CH2 SÞ2 2 þ 4H2 O

Kf, 1 ¼

½ðVO2 Þ2 ðOCH2 CH2 SÞ2 2 -  ½H2 VO4- 2 ½HOCH2 CH2 SH2

ð10Þ ð11Þ

A minor component, complex 2, was observed at high vanadate concentrations. As the solution ionic strength increased, the 51V signal intensities for complex 2 (-385 ppm), V4 (-577 ppm), and V5 (-586 ppm) increased, whereas the signal intensities for complex 1 (-362 ppm), V1 (-542 ppm), and V2 (-566 ppm) decreased (data not shown). The greater sensitivity of complex 2 to ionic strength is consistent with this species being a higher oligomer than complex 1. The low concentration of complex 2 at modest vanadate concentrations required that this species be studied at high ionic strength.64 The 51 V chemical shifts for both complexes at -362 and -385 ppm do not change from pH 7.0 to 10.5 consistent with no change in the charge of either complex. Solutions containing vanadate (50-150 mM) and 2-mercaptoethanol (50-150 mM) in the presence of 3.0 M KCl at pH 8.4 (primarily) were used to determine the stoichiometry and formation constant of complex 2. The expectation that 2 is an oligomeric species was confirmed by the linear relationship between [2] and [1]2 shown in Figure 3b. Complex 2 has a stoichiometry of 4:4 (eq 12, 13). Other possible stoichiometries were ruled out (see Supporting Information, Figure S3). Experimental difficulties with variation of the solution pH at high vanadate concentrations limited the number of measurements made at other pH values. Using the pKa values of vanadate and 2-mercaptoethanol in 3.0 M KCl, the Hþ-independent formation constant for the 4:4 species (eq 12, 13) was calculated to be (3.9 ( 0.1) 3 1016 M-7. 4H2 VO4 - þ 4HOCH2 CH2 SH a ½ðVO2 Þ4 ðOCH2 CH2 SÞ4 4 þ 8H2 O

Kf, 2 ¼

½ðVO2 Þ4 ðOCH2 CH2 SÞ4 4 -  ½H2 VO4- 4 ½HOCH2 CH2 SH4

ð12Þ ð13Þ

The structure of (NEt4)2[(VO2)2(OCH2CH2S)2]26 demonstrates an analogy with vanadate-1,2-diolate structures.53-55,75-79 Higher oligomeric species akin to complex 2 are unprecedented in aqueous solution, but have been reported in studies in organic solvents.80-82 Solution Structures for the Vanadium(V) Complexes with 2-Mercaptoethanol. Using the Coordination Induced Shift (CIS, where Δδ = δcomplex - δfree ligand)83,84 in 13C and 1 H NMR spectroscopic studies, we deduced the solution structure for Complex 1. The 1H and 13C NMR data for complexes 1 and 2, free ligand, and CIS values at pH 8.9 are compiled in Table 2. The CIS value for the carbon adjacent to the oxygen is 15-19 ppm is in the range of CIS values commonly observed for carbon adjacent to bridging oxygen atoms.83,84 The CIS shift for the carbon adjacent to the sulfur is 8.3 ppm. Since limited information is available for CIS shifts of vanadium(V)-thiol complexes, trithiatren [(HSCH2CH2)3N]85 and its vanadium(V) complex [VO(SCH2CH2)3N]31 were synthesized to determine its 13C NMR CIS value. The CIS value for the carbon atoms adjacent to the sulfur in [VO(SCH2CH2)3N] in CD3CN was 14.5 ppm. Unfortunately, this species was not sufficiently soluble in aqueous solution to obtain the CIS value. However, solvent variations in CIS values for related oxygen derivatives show no more than 2-4 ppm variation because of changes in solvent.83 Furthermore, a 1:1 complex of chromate and glutathione (GSH) has a CIS value, for the carbon atoms adjacent to sulfur, of 11.0 ppm.19 Since the CIS value in complex 1 is 8.3 ppm the V-S bond in this system is weaker than those described above and suggest that the sulfur is not bridged. (75) Gresser, M. J.; Tracey, A. S. J. Am. Chem. Soc. 1986, 108, 1935–1939. (76) Tracey, A. S.; Gresser, M. J. Inorg. Chem. 1988, 27, 2695–2702. (77) Geraldes, C. F. G. C.; Castro, M. M. C. A. J. Inorg. Biochem. 1989, 35, 79–93. (78) Tracey, A. S.; Leonlai, C. H. Inorg. Chem. 1991, 30, 3200–3204. (79) Richter, J.; Rehder, D. Z. Naturforsch., B: Chem. Sci. 1991, 46, 1613– 1620. (80) Heinrich, D. D.; Folting, K.; Streib, W. E.; Huffman, J. C.; Christou, G. Chem. Commun. 1989, 1411–1413. (81) Rehder, D.; Priebsch, W.; Vonoeynhausen, M. Angew. Chem., Int. Ed. Engl. 1989, 28, 1221–1222. (82) Cavaco, I.; Pessoa, J. C.; Duarte, M. T.; Matias, P. M.; Henriques, R. T. Polyhedron 1993, 12, 1231–1237. (83) Crans, D. C.; Chen, H. J.; Anderson, O. P.; Miller, M. M. J. Am. Chem. Soc. 1993, 115, 6769–6776. (84) Crans, D. C.; Felty, R. A.; Chen, H. J.; Eckert, H.; Das, N. Inorg. Chem. 1994, 33, 2427–2438. (85) Barbaro, P.; Bianchini, C.; Scapacci, G.; Masi, D.; Zanello, P. Inorg. Chem. 1994, 33, 3180–3186.

Article

Inorganic Chemistry, Vol. 49, No. 9, 2010

4251

Table 2. NMR Data for 2-Mercaptoethanol, 2-Hydroxyethyl Disulfide and Complexes 1 and 2 1

compound HOCH2CH2SH (HOCH2CH2S)2 1 2 a

51

V ppm

-362 -385

17

O ppm

988 802

13

H ppm (CIS)a

C ppm (CIS)a

H1

H2

C1

C2

3.70 3.97(0.27) 4.31(0.61) 4.23(0.53)

2.68 2.90(0.22) 3.05(0.37) 2.97(0.29)

66.1 62.1(-4.1) 85.3(19.2) 83.1(17.0)

28.8 42.8(14.0) 37.1(8.3) 36.1(7.3)

Numbers in parentheses indicate CIS (CIS = δcomplex - δfree ligand).83,84

Figure 4. Solution structure for the 2:2 vanadium complex with 2mercaptoethanol, 1, and two likely structures for the 4:4 vanadium complex with 2-mercaptoethanol, 2. Only one of several possible stereoisomers is shown.

However, a CIS of 8.3 ppm is significant and consistent with formation of a covalent bond. These deductions yield a structure which is identical to that observed in the solid state26 (Figure 4, 1), and support the report that [(VO2)2(OCH2CH2S)2]2- exists in solution. Oxygen-17 NMR spectroscopy has been successfully used to distinguish between terminal (non-bridging) oxygen atoms and bridging oxygen atoms in vanadates86,87 and in molybdates,88 as well as to provide structural information in vanadate-1,2-diolates and other complexes.53,89 The 51V and 17O NMR spectra of complex 1 were obtained from a concentrated solution of vanadate (300 mM) and 2-mercaptoethanol (600 mM) in 3.5 atom % 17O enriched water at pH 8.6. The 51V NMR spectrum (Supporting Information, Figure S4 (top)) shows one major peak for complex 1 (97.8%), one minor peak for complex 2 (1.2%), and two minor resonances (1.0%) attributed to vanadate monomer and oligomers. The major peak in 17O NMR spectrum (Supporting Information, Figure S4 (bottom)) with a line width of 407 Hz and a chemical shift of 988 ppm, relative to H2[17O], accounts for 98.2% of the 17O signals observed (exclusive of H2[17O]) in the region from -100 ppm to þ1500 ppm. No signals are observed for H2VO4- or HVO42-/VO43which have 17O shifts of 573 and 565 ppm.87 We assign the resonance at 988 ppm to the oxovanadium oxygen atoms of complex 1. Integration of a minor peak observed at 802 ppm resulted an intensity of less than 2%, consistent with this signal arising from a different species such as complex 2 and not from a different type of oxygen atom in (86) Day, V. W.; Klemperer, W. G.; Maltbie, D. J. J. Am. Chem. Soc. 1987, 109, 2991–3002. (87) Heath, E.; Howarth, O. W. J. Chem. Soc., Dalton Trans. 1981, 1105– 1110. (88) Filowitz, M.; Klemperer, W. G.; Messerle, L.; Shum, W. J. Am. Chem. Soc. 1976, 98, 2345–2346. (89) Crans, D. C.; Shin, P. K. J. Am. Chem. Soc. 1994, 116, 1305–1315.

complex 1. The similarity of the 17O NMR spectra of the vanadate-2-mercaptoethanol (988 ppm) and the vanadate-ethylene glycol complexes (1004 ppm)53 support the solution structure being the same as that found in the solid state.26 The following consideration for complex 2 supports the proposed two structures 2a and 2b shown in Figure 4 with the caveat that the vanadium atoms in structure 2b are rapidly exchanging as reported previously.92 The 13C CIS of 17.0 ppm for the carbon adjacent to the oxygen suggests the alkoxide oxygen to be bridging. The CIS of 7.3 ppm for the carbon adjacent to the sulfur indicates a covalent but non-bridging V-S bond. The 1H NMR data support these conclusions (see Table 2). The 51V chemical shift of complex 2 (-385 ppm) is only 23 ppm upfield of complex 1 (-362 ppm) consistent with one oxygen and one sulfur atom of the ligand coordinated to each vanadium atom in complex 2. The 17O NMR signal supports either a rapidly exchanging tetrameric structure such as 2b or a structure in which the VdO groups are identical or sufficiently similar to have indistinguishable 17O chemical shifts such as 2a. Although there are many structural possibilities for complex 2, the stoichiometry and the similarities in the CIS values between the two complexes rule out most of these possibilities, leaving structures 2a and 2b as likely candidates. Structures 2a and 2b are related to various tetra-nuclear vanadium alkoxides giving precedence for these types of structures as well as other alkoxides with the [V(OR)OV]-unit.80-82,90-94 Vanadium(V) Complexes with 2-Mercaptoethanol in Mixed Aqueous-Acetonitrile Solutions. Since the isolated complex (NEt4)2[(VO2)2(OCH2CH2S)2] dissolved in CH3CN has a chemical shift of -385 ppm in between complexes 1 and 2, spectra were recorded at different D2O/CH3CN ratios to reconcile the observation of only one complex in CH3CN.26 A series of 51V spectra of solutions containing 50 mM vanadate and 50 mM 2-mercaptoethanol were recorded in D2O and CH3CN (from 0 to 87 wt %) and are shown in Figure 5. The peaks and intensities for complexes 1 and 2 in D2O (0 wt % CH3CN) are perturbed when CH3CN is added. Increasing the concentration of CH3CN from 0 to 87 wt % results in ∼10 ppm shift of the major signal for complex 1.26 The signals from complex 2 and vanadate oligomers (90) Drouza, C.; Keramidas, A. D. Inorg. Chem. 2008, 47, 7211–7224. (91) Crans, D. C.; Jiang, F. L.; Chen, J.; Anderson, O. P.; Miller, M. M. Inorg. Chem. 1997, 36, 1038–1047. (92) Crans, D. C.; Marshman, R. W.; Gottlieb, M. S.; Anderson, O. P.; Miller, M. M. Inorg. Chem. 1992, 31, 4939–4949. (93) Mikuriya, M.; Kotera, T.; Adachi, F.; Bandow, S. Chem. Lett. 1993, 945–948. (94) Reis, D. M.; Westrup, K. C. M.; Nunes, G. G.; Barison, A.; Ribeiro, R. R.; de Sa, E. L.; Hughes, D. L.; Soares, J. F. J. Braz. Chem. Soc. 2009, 20, 613–626.

4252 Inorganic Chemistry, Vol. 49, No. 9, 2010

Crans et al.

Figure 5. 51V NMR spectra of solutions containing KVO3 (50 mM) and 2-mercaptoethanol (50 mM) at pH 8.25 ( 0.05. CH3CN is added up to 87 wt %, and spectra at 0, 68, 75, 81, 84, and 87 wt % are shown.

decrease in number and intensity upon addition of CH3CN. Complex 2 did not shift with increasing concentration of CH3CN, but its intensity decreased. These studies confirm that complex 1 forms in both aqueous and CH3CN as reported previously,26 but that complex 2 is only observed in our studies. Complex 2 is not observed in aqueous solutions of vanadate and 2-mercaptoethanol when vanadate concentrations are below 8 mM. When complex 1 is formed from solid (NEt4)2[(VO2)2(OCH2CH2S)2] its association, dissociation, and/or hydrolysis is slow, preventing complex 2 formation in these solutions. This hypothesis was confirmed using VT 51V NMR spectroscopy in an aqueous solution of 50 mM of vanadate and 50 mM of 2-mercaptoethanol (pH 8.2) where complexes 1, 2, and vanadate were all present. At increasing temperatures the signals for complex 1 and 2 did not coalesce, consistent with no dissociation or association between the two complexes and vanadate. However, as the temperature was increased above 348 K, the concentration of complex 2 decreased below the detection limit. These studies document the kinetic inertness of complex 1, and low concentration and narrow temperature range over which complex 2 exists. Characterization of the Vanadium(IV)-2-Mercaptoethanol Complex: Stoichiometry and Formation Constant. Mixing a solution of VOSO4 with 2-mercaptoethanol in the neutral or slightly alkaline pH range generates an EPR spectrum of a new 2-mercaptoethanol complex (complex 3; parameters: g0, 1.976 and A0, 81  10-4 cm-1). Solutions of VOSO4 in this pH range are grayish-green and turn bright green upon addition of 2-mercaptoethanol. In Figure 6 we show EPR spectra of 10 mM VOSO4 solutions with increasing concentrations (up to 200 mM) of 2-mercaptoethanol at pH 9.6 in borate buffer. In these solutions increasing concentrations of complex 3 are observed. Formation of 3 is pH dependent as shown in Figure 7 from pH 3.2 to 13.4. At acidic pH only [VO(H2O)5]2þ is observed (g0, 1.968 and A0, 107  10-4 cm-1); however, as the pH is increased above 7 the 2-mercaptoethanol complex emerges. A single species, complex 3, is shown at pH 10.4 in Figure 7.

Figure 6. EPR spectra recorded in solutions containing VOSO4 (10 mM) and increasing concentrations of 2-mercaptoethanol (from 0 to 200 mM) in 0.60 M borate buffer (pH 9.6).

Figure 7. Normalized EPR spectra recorded in solutions with VOSO4 (10 mM) and 2-mercaptoethanol (200 mM) at pH 3.2 to 13.4.

Complex 3 forms in the pH range near the pKa value of 2-mercaptoethanol (9.7) consistent with the deprotonated 2-mercaptoethanol being the coordinating ligand. In Figure 8 the concentrations of complex 3 and [VO(OH)3(H2O)2]- (g0, 1.973 and A0, 82.9  10-4 cm-1)66 are shown in solutions containing a total of 10 mM VOSO4

Article

Inorganic Chemistry, Vol. 49, No. 9, 2010

Figure 8. Concentration of complex 3 and [VO(OH)3]- ([VO(OH)3(H2O)2]-) are shown as a function of pH. Data points were obtained by quantitative EPR spectroscopy in solutions with VOSO4 (10 mM) and 2-mercaptoethanol (200 mM). No complex signal is observed outside the pH region 8-14. The smooth line combining the experimental points for [3] and [VO(OH)3]- is calculated using the stability constants95 and the reaction VO2þ þ 2 HOCH2CH2S- / [VO(OCH2CH2S)2]2- þ 2 Hþ.

and 200 mM of 2-mercaptoethanol over the pH range of 7 to 14. The complex is most stable in the pH range 9.5 to 11.5 (Figure 8). Since high concentrations of [VO(OH)3(H2O)2]- are observed above pH 10.8, evaluation of the EPR spectrum of 3 is best carried out from pH 9.2 to 10.8 where the spectrum is not complicated by the signal of [VO(OH)3(H2O)2]-. We therefore chose to study the complex at pH 9.6 in borate buffer. The maximum concentration of 3 was found to be 4.88 mM although a total of 10.0 mM of VOSO4 was added to the solution. The additional species existing in solution are EPR silent vanadium(IV) (most likely a dimer [(VO)2(OH)5]-) since, up to a pH of 8.5, less than 0.5% of the vanadium(IV) was oxidized within 2 h of addition of 2-mercaptoethanol, as determined by quantitative 51 V NMR spectroscopy. After 20 h at pH 8.5 the vanadium(V) complex content has increased to about 2%. At a higher pH, such as 10.3, 4% of the vanadium(IV) was oxidized within 2 h and 15% within 20 h; however, a significant fraction of the vanadium added to the solutions is still present as EPR silent vanadium(IV) complexes. As the pH was increased to 12.3 and 13.3, the fraction of oxidized vanadium(IV) increased further to 11% and 30% within 2 h. In the experiments shown in Figure 8, the amount of oxidized vanadium is negligible since the EPR were recorded within 30 min of preparation. pVO2þ þ rHL - a ½ðVOÞp ðLÞr ðHÞðr - qÞ ð2p - r - qÞ þ qHþ ð14Þ ð2p - r - qÞ

Kf, 3 ¼

½ðVOÞp ðLÞr ðHÞðr - qÞ

½VO2þ p ½HL - r

½Hþ q

ð15Þ

The stoichiometry of the complex was determined by quantitative EPR spectroscopy for freshly prepared solutions adding 10.0 mM VOSO4 and from 20 to 200 mM 2-mercaptoethanol in 0.40 M borate buffer at pH 9.6. Equations 14 and 15 describe a general reaction of HL(HL- = HOCH2CH2S-) with [VO(H2O)5]2þ. The concentration of free [VO(H2O)5]2þ was calculated from the

4253

Figure 9. Plot of complex [3] in borate buffer at pH 9.6 as a function of [VO(H2O)52þ][HL-]2 (here HL- = HOCH2CH2S-). The linear relationship indicates a 1:2 complex stoichiometry for 3. Other possibilities result in non-linear plots (Supporting Information, Figure S5).

concentration of [VO(OH)3(H2O)2]- assuming three major species are present in the pH region of interest: complex 3, [(VO)2(OH)5]-, and [VO(OH)3(H2O)2]-. Plotting [3] as a product of [VO(H2O)52þ][HOCH2CH2S-]2 gives a linear plot (Figure 9), which defines the p and r values (eq 14) as 1 and 2, respectively. Other p/q ratios such as 1:1, 2:2, and 1:3 were ruled out (Supporting Information, Figure S5). Only with q = 2 did the simulated curve of [3] as a function of pH give a maximum near pH 10 and model the complex concentration found experimentally. In Figure 8 the calculated curve is superimposed on the experimental points obtained from EPR studies. On the basis of these considerations, the complex has a (2-) charge, and its formation constant defined by eq 15 (p = 1, q = 2, and r = 2) is equal to 5.1 ((0.2)  10-7. Modeling studies using species with various p/r ratios (such as 1:1, 1:2, 2:1, 2:2, 2:3, and 2:4) and q ranging from -2 to 2 did not yield as good a fit to the experimental data. These calculations are consistent with complex 3 being [VO(SCH2CH2O)2]2- in solution. Proposed Solution Structures for the Vanadium(IV) Complexes with 2-Mercaptoethanol. We measured the ambient (300 K) and low (140 K) temperature EPR spectra of 10 mM vanadyl sulfate and 200 mM 2-mercaptoethanol at pH 9.6. The simulation of the EPR spectra at ambient temperature (Figure 7, pH 10.4) led to a g0 of 1.976 and an A0 of 81.0  10-4 cm-1. From the low temperature spectrum (not shown), the anisotropic g factors and hyperfine splitting constants were obtained as g||, 1.967; g^, 1.980; A||, 146  10-4 cm-1; A^, 41.0  10-4 cm-1. The parameters for [VO(H2O)5]2þ were g0 of 1.968 and an A0 of 107  10-4 cm-1 at room temperature (Figure 7, pH 3.2), and g|| of 1.933, g^ of 1.985, A|| of 182  10-4 cm-1, and A^ of 68.4  10-4 cm-1 at 120 K (spectrum not shown) which are in agreement with previous reports.66,96,97 The parameters for [VO(OH)3(H2O)2]- species were g0 of 1.973 and A0 of 82.9  10-4 cm-1 at ambient temperature (Figure 7, pH 13.4), and g|| of 1.951, g^ of 1.978, A|| of (95) The parameters used in the simulation are: log β [VO(OH)3-] = -19.25; log β [(VO)2(OH)5-] = -22.8; log β [VO(OCH2CH2S)22-] = -5.68; pKa (HOCH2CH2SH) = 10.3. (96) Boucher, L. J.; Tynan, E. C.; Yen, T. F. In Spectral properties of oxovanadium(IV) complexes; Yen, T. F., Ed.; Plenum: New York, 1969; pp 111-130. (97) Smith, T. S.; LoBrutto, R.; Pecoraro, V. L. Coord. Chem. Rev. 2002, 228, 1–18.

4254 Inorganic Chemistry, Vol. 49, No. 9, 2010

Crans et al.

Figure 10. Four possible solution structures for complex 3.

16110-4 cm-1, and A^ of 46.9  10-4 cm-1 at low temperature (spectrum not shown). Four possible structural isomers can be proposed for complex 3, Figure 10. Since the equatorial ligands provide additive contributions to the A|| an estimation for each structure can be calculated.98 The partial components in cm-1 for H2O (45.7  10-4), aliphatic alkoxide (35.3  10-4), and aliphatic thiolate (31.9  10-4), yields a A|| value of 145  10-4 cm-1 for A1, 148  10-4 cm-1 for A2, and 134  10-4 cm-1 for both B1 and B2, respectively. The observed A|| value of 146  10-4 cm-1 is consistent with complex 3 being the six-coordinate species A1 or A2 containing a cis-H2O. However, examples of both cis-99 and trans-adducts57 have been reported. A detailed description of the vanadium(IV)- glutathione system confirms multiple species with a polydentate glutathione bound to the oxovanadium center.100 Isolation and structural characterization of a range of oxo and non-oxo vanadium-sulfur complexes support the types of structures proposed in this system.2,5-11 UV-vis Spectroscopic Characterization of the Reduction of Vanadate to Vanadyl by 2-Mercaptoethanol. The reduction of vanadate by 2-mercaptoethanol was strongly dependent on solution pH, solute concentration, and the presence of molecular oxygen. The visible spectrum 5 min after preparation of the yellow solution containing 10 mM vanadate and 200 mM 2-mercaptoethanol at pH 8.9 is shown in Figure 11. This spectrum is attributed to the 2:2 complex, since the concentration of the 4:4 complex is negligible under these conditions. After 5 h the solution color had changed to green, with visible spectrum absorption maxima at 558 and 625 nm (Figure 11, pH = 8.9, 5 h). A solution of vanadyl sulfate (10 mM) and 2-mercaptoethanol (200 mM) gives the same absorption spectrum at pH 8.9 (data not shown), consistent with attributing the green color to a vanadium(IV)-2-mercaptoethanol complex. When the pH of this solution was decreased below pH 4, a blue solution and the visible spectrum characteristic of vanadyl cation resulted. Identical spectra are obtained from solutions of vanadyl sulfate (10 mM) in the absence or presence of 2-mercaptoethanol (200 mM) at pH 3.0 (Figure 11, pH 3.0, 5 min). The maximum at 768 nm and the shoulder at 628 nm obtained from a solution of sodium vanadate in the presence of excess thiol indicate the formation of a (98) Rehder, D. In Bioinorganic Vanadium Chemistry; Rehder, D., Ed.; John Wiley & Sons: Chichester, England, 2008; pp 67-75. (99) Hanson, G. R.; Sun, Y.; Orvig, C. Inorg. Chem. 1996, 35, 6507–6512. (100) Pessoa, J. C.; Tomaz, I.; Kiss, T.; Kiss, E.; Buglyo, P. J. Biol. Inorg. Chem. 2002, 7, 225–240.

Figure 11. Absorption spectra of the vanadate- and vanadyl-2-mercaptoethanol complexes prepared under anaerobic conditions from vanadate (20 mM) and 2-mercaptoethanol (200 mM) at pH 8.9 at different times after the solution preparation: (a) 5 min; (b) 5 h; and (c) at pH 3.0 (not time dependent).

vanadium(IV) complex. A similar spectrum was also observed from a solution containing 2-mercaptoethanol, and the microcrystalline blue-black vanadium(IV) oxide precipitated from solutions at high pH. On the basis of these results, the formation of vanadium(IV)-2-mercaptoethanol complex in vanadate solutions follows the reduction of vanadate by 2-mercaptoethanol. The process is summarized in eq 16 and 17. 2HOCH2 CH2 SH þ 2H2 VO4 - þ 6Hþ a ðHOCH2 CH2 SÞ2 þ 2VO2þ þ 6H2 O

ð16Þ

VO2þ þ 2HOCH2 CH2 S - a ½VOðSCH2 CH2 OÞ2 2 - þ 2Hþ

ð17Þ The reduction of vanadate by 2-mercaptoethanol uses 3 equiv of Hþ consistent with the facile reduction in acidic media. At neutral pH or in the presence of excess vanadate, however, the gray suspension of soluble and insoluble vanadyl hydroxides will form and eventually generate the blue-black suspension of V2O4.49 An excess of 2-mercaptoethanol will complex vanadyl cation in alkaline solution where sufficiently high concentrations of thiolate exist and can remain in solution for some time. Observations by UV-vis and/or EPR spectroscopy have been reported with related systems.12-15,101 In our hands vanadate(V) (20 mM) and glutathione (200 mM) at pH 8.9 produced a green solution in agreement with the reports that vanadate is reduced to vanadium(IV) in the presence of glutathione.16,101 The solution of vanadate and glutathione yielded a similar UV-visible spectrum and EPR spectrum (data not shown) to the 2-mercaptoethanol complex. Interestingly the complex formed from cysteine and vanadyl cation13 is purple, as is the complex between cysteine methyl ester and vanadyl cation.102 Presumably this color is due to formation of a VO(S2N2)-type complex. (101) Macara, I. G.; Kustin, K.; Cantley, L. C. Biochim. Biophys. Acta 1980, 629, 95–106. (102) Sakurai, H.; Hamada, Y.; Shimomura, S.; Yamashita, S.; Ishizu, K. Inorg. Chim. Acta 1980, 46, L119–L120.

Article

Comparing the Stability of Vanadium(IV) and (V) 2-Mercaptoethanol Complexes. Given the fact that 2-mercaptoethanol forms complexes with both vanadium(V) and vanadium(IV), the question of what kind of vanadium species are present at neutral pH depends on the aqueous V(V) and V(IV) speciation chemistry. Unique species are present in both oxidation states, and more complex 1 is present than complex 3 based on the distribution diagrams (Figures 2 and 8). Calculating concentrations of vanadium species for two hypothetical cases at total vanadium concentrations of 1 mM and 1 μM at pH 7 confirmed that [1].[3],103 i.e., the amount of vanadium(V) bound to the ligand is greater than that of vanadium(IV) from 106-fold (at 1 mM) to 500-fold (at 1 μM). The formation of the V(V)2-mercaptoethanolate and significantly less stable V(IV)-2mercaptoethanolate complexes suggests that vanadate can coexist in solution with 2-mercaptoethanol because of the low vanadate oxidation potential. Our calculations carried out for a solution containing 1.0 mM vanadium(V), 1.0 mM vanadium(IV), 10 mM 2-mercaptoethanol, and 0.1 mM 2-hydroxyethyl disulfide show that the vanadium(V)/ vanadium(IV) reduction potential reaches 2-mercaptoethanol/2-hydroxyethyl disulfide oxidation potential at pH 11.5 ((0.5). Below pH 11.5 vanadate oxidizes thiol to disulfide. Rate of Complex 1 Reduction under Anaerobic Conditions. Having documented that 2-mercaptoethanol reduces vanadium(V) to vanadium(IV), it is an important question to determine how fast complex 1 decomposes. The redox reaction proceeds readily at low pH and high solute concentrations with consumption of 3 equiv of Hþ (eq 16). Because of Hþ consumption, kinetic studies should be carried out in buffered solution. To minimize thiol oxidation by oxygen (which is rapid in basic solution)104 the reaction solution was nitrogen purged throughout the experiment. Borate was selected as buffer because of its buffering capacity (pKa 9.24) at pH 8.9 and its lack of 1H NMR signals. Although no new signal is observed in the 51 V NMR spectra of vanadate solutions containing borate, an increase in the linewidths of the V1 and V2 signals is indicative of an interaction. This interaction is likely to be weak, as previously observed for phosphate,105 acetate,106 chromate,107 and other anions.64 Given the lack of kinetic (103) The calculation for distribution of vanadium at pH 7.0 carried out for two cases (i) [V(V)]TOT =[V(IV)]TOT = 1.0 mM, [2-mercaptoethanol]TOT = 10 mM, and (ii) [V(V)]TOT = [V(IV)]TOT = 1.0 μM, [2-mercaptoethanol]TOT = 10 mM (Supporting Information, Table 1S) yielded the following results: (a) For the 1.0 mM solution, vanadium(V) species are 75 mol % complex 1, 20 mol % V1, 3.4 mol % V2, and 1.2 % V4. For the 1 μM solution the major species are 98 mol % V1 and 2.0 mol % complex 1. (b) In the 1 mM vanadium(IV) solutions the solubility is limited by the formation of insoluble [VO(OH)2]. The major vanadium(IV) species at 1.0 mM are 28 mol % [(VO)2(OH)5]- and 0.21 mol % [VO(OH)3]-. At 1 μM the major species are 85 mol % [(VO)2(OH)5]and 12 % [VO(OH)3]-. Complex 3 is only a minor species, present at the trace levels 7.6  10-5 mol % and 4.2  10-3 mol % in the 1 mM and 1 μM solutions, respectively. All percent values in this section represent molar percent of vanadium in oxidation states (IV) and (V). (104) Oswald, A. A.; Wallace, T. J. In The Chemistry of Organic Sulfur Compounds; Kharasch, N., Meyers, C. Y., Eds.; Pergamon: New York, 1966; Vol. 2, pp 205-232. (105) Gresser, M. J.; Tracey, A. S.; Parkinson, K. M. J. Am. Chem. Soc. 1986, 108, 6229–6234. (106) Tracey, A. S.; Li, H. L.; Gresser, M. J. Inorg. Chem. 1990, 29, 2267– 2271. (107) Howarth, O. W.; Winter, D. J. Chem. Soc., Dalton Trans. 1991, 399–402.

Inorganic Chemistry, Vol. 49, No. 9, 2010

4255

Figure 12. [1] and [2], [HOCH2CH2SH], and [(HOCH2CH2S)2] during the first 3 h of reaction. The initial solution were added vanadate (200 mM), HOCH2CH2SH (200 mM), KCl (0.4 M), and borate buffer (0.4 M) at pH 8.95.

data on these types of reactions, even limited information on the reactivity of the system is valuable. A series of 1H and 51V NMR spectra were recorded over 3 h on solutions 0.40 M borate and 0.40 M KCl with varying concentrations of vanadate and 2-mercaptoethanol. The vanadium(V) species rapidly equilibrate108 with each other so that integrations of the 1H and 51V NMR spectra give the concentrations of each vanadium(V) species. Figure 12 shows the typical changes in concentration of complexes 1 and 2, 2-mercaptoethanol, and 2-hydroxyethyl disulfide as a function of time. While the concentration of the disulfide increases, the concentrations of 1, 2, and the thiol decrease, albeit at different rates. In addition, the concentration of V1 also decreases. On the basis of initial reaction rates, we obtained a reaction half-life time t1/2 > 16 h in concentrated solutions (200 mM of each reactant) and t1/2 > 87 h (25 mM of each reactant). Under anaerobic conditions complex 1 decomposes only slowly which is in contrast to the decomposition observed in the presence of oxygen (data not shown). Further mechanistic details of the reaction are beyond the scope of this paper. Conclusion This work confirms the fact that vanadate can form complexes with thiols without undergoing reduction even though the same thiol under other conditions will induce redox chemistry. Although complex formation suppresses redox chemistry, these studies define conditions under which the reactants do undergo redox chemistry. In the specific case of 2-mercaptoethanol vanadate mainly forms a stable complex (1, δ(51V) -362 ppm) of 2:2 stoichiometry with 2-mercaptoethanol at neutral or alkaline pH. A minor 4:4 complex (2, δ(51V) -385 ppm) can be observed at high concentrations of vanadate and 2-mercaptoethanol. Vanadate is reduced in the course of oxidizing 2-mercaptoethanol to the disulfide, and rates are highest at low pH and high thiol concentrations. The reduction is sufficiently slow for observation of the coordination chemistry. Complex 3 that forms between aqueous vanadium(IV) and 2-mercaptoethanol has a 1:2 stoichiometry and is sufficiently stable to be observable at high pH and high 2-mercaptoethanol (108) Crans, C. D.; Rithner, C. D.; Theisen, L. A. J. Am. Chem. Soc. 1990, 112, 2901–2908.

4256 Inorganic Chemistry, Vol. 49, No. 9, 2010 concentration. The structure of this complex was deduced from its EPR parameters. Both bis(2-mercaptoethanolate) complexes (V(IV) and V(V)) contain deprotonated ligands. However, the stoichiometries of the complexes in the two oxidation states are different. The amounts of vanadium(IV) and (V) complexes were measured in the same solution which is important to the evaluation of the action of vanadium complexes in biological systems. Issues regarding literature reports and seeming discrepancies concerning whether or not a particular thiol will or will not reduce vanadate are clarified. We demonstrate that reactions, in which the components’ standard reduction potentials are similar, are strictly governed by solution pH, solute concentration, and the stability of a potential complex. In this work we show that the high stability of a vanadium(V) complex can result in its formation, an observation that flies in the face of the common misconception that vanadium(IV) complexes are more stable.11,107-109 The work described in this paper explains the apparent controversy regarding whether or not vanadium(V) (109) Rehder, D. Inorg. Chem. Commun. 2003, 6, 604–617.

Crans et al. compounds oxidize thiols under physiological conditions. The redox behavior of vanadium complexes are likely to be important to their insulin-enhancing properties,43,109-113 and this study provides some guidelines when complexes are reduced. Acknowledgment. D.C.C. thanks the General Medical Institute at the National Institutes of Health (GM40525). We also thank Drs. Feibo Xin and Sandra M. Eaton for technical assistance during the course of this work. Supporting Information Available: Further details are given in Table 1S and Figures S1-S6. This material is available free of charge via the Internet at http://pubs.acs.org. (110) Sakurai, H.; Yoshikawa, Y.; Yasui, H. Chem. Soc. Rev. 2008, 37, 2383–2392. (111) Crans, D. C.; Yang, L. Q.; Jakusch, T.; Kiss, T. Inorg. Chem. 2000, 39, 4409–4416. (112) Buglyo, P.; Crans, D. C.; Nagy, E. M.; Lindo, R. L.; Yang, L. Q.; Smee, J. J.; Jin, W. Z.; Chi, L. H.; Godzala, M. E.; Willsky, G. R. Inorg. Chem. 2005, 44, 5416–5427. (113) Li, M.; Ding, W.; Smee, J. J.; Baruah, B.; Willsky, G. R.; Crans, D. C. Biometals 2009, 22, 895–905.